This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Letter pubs.acs.org/NanoLett

Unimolecular Submersible Nanomachines. Synthesis, Actuation, and Monitoring Víctor García-López,† Pinn-Tsong Chiang,† Fang Chen,‡ Gedeng Ruan,† Angel A. Martí,*,† Anatoly B. Kolomeisky,*,†,§ Gufeng Wang,*,‡ and James M. Tour*,†,∥ †

Department of Chemistry, §Department of Chemical and Biomolecular Engineering and Center for Theoretical Biological Physics, Department of Materials Science and NanoEngineering, Rice University, Houston, Texas 77005, United States ‡ Department of Chemistry, North Carolina State University, Raleigh, North Carolina 27695, United States ∥

S Supporting Information *

ABSTRACT: Unimolecular submersible nanomachines (USNs) bearing light-driven motors and fluorophores are synthesized. NMR experiments demonstrate that the rotation of the motor is not quenched by the fluorophore and that the motor behaves in the same manner as the corresponding motor without attached fluorophores. No photo or thermal decomposition is observed. Through careful design of control molecules with no motor and with a slow motor, we found using single molecule fluorescence correlation spectroscopy that only the molecules with fast rotating speed (MHz range) show an enhancement in diffusion by 26% when the motor is fully activated by UV light. This suggests that the USN molecules give ∼9 nm steps upon each motor actuation. A non-unidirectional rotating motor also results in a smaller, 10%, increase in diffusion. This study gives new insight into the light actuation of motorized molecules in solution. KEYWORDS: Unimolecular submersible nanomachines, light-driven motor, diffusion coefficient, fluorophores nspired by the “bottom up” approach1−3 used by nature to build functional macroscopic entities using nanoscopic buildings blocks, synthetic chemists have designed a variety of molecular machines and nanovehicles such as nanoscale motors, switches, turnstiles, barrows, shuttles, and nanocars.4 Specifically, we have used scanning tunneling microscopy (STM)5−7 and single molecule fluorescence microscopy (SMFM)8−12 to track nanocars on surfaces. However, these imaging methods cannot be directly applied to unimolecular nanomachines in solution because they drift quickly out of focus in three-dimensional (3D) environments, thus producing trajectories that are too short to determine accurate diffusion coefficients. As biological processes take place in solution, the development of nanomachines that are able to enhance their diffusion and perform work in that phase is of great interest. This has led to the development of self-propelled nanowires,13,14 microrockets,15 Janus-particle motors,16,17 enzymatic motors,18,19 and mineral micropumps20 powered by chemical reactions through self-electrophoretic mechanisms, bubble propulsion, or difusioosmosis. However, most of those micromachines use or generate toxic chemicals that are inappropriate for in vivo applications. To address the disadvantage of using toxic chemicals, cleaner systems that convert photonic energy to translational motion have been developed. Silver chloride particles21 and TiO2 micromotors22,23 are some examples of

I

© 2015 American Chemical Society

micromachines able to move in solution under UV light illumination via a self-diffusiophoresis mechanism. All of the micromachines mentioned above range from hundreds of nanometers to micrometers in size. At present, there are only two examples of catalytically driven unimolecular nanomachines (99.8%, suggesting that the diffusion of USN-4 is enhanced in the presence of UV activation. The smaller enhancement in the diffusion of UNS-4 could result from two possible reasons: (1) the step size of the molecule is smaller or (2) the rotation speed is slower. The rate at which the nondirectional rotor in USN-4 moves is unknown. We know that if the stereogenic center-appended methyl

absolute diffusion coefficient of molecules depends on experimentally adjustable parameters such as laser beam waist-size. Such parameters may vary slightly from time to time, introducing errors to the measurements.41 To minimize these systematic errors, the FCS experiments with and without UV excitation were always collected in pairs using the same solution and at the same collection spot. Hence, the only contrast was with or without UV light illumination. The sequence of collection has no observable effect on the diffusion coefficient measurements. In the absence of UV light activation, USN-1 diffuses freely in bulk solutions. The autocorrelation function (ACF) can be satisfactorily fitted with the 3D diffusion model (Figures S2 and S3). The diffusion coefficient (D) of USN-1 was 0.92 ± 0.07 × 10−10 m2·s−1 (95% confidence interval from Student’s t test) from repeated measurements on different days and for different samples. This D is on the same order of magnitude for other small molecules in ACN.42 When the UV light was turned on, the diffusion of USN-1 becomes faster. This can be viewed from the ACF decays. Figure 4A shows the normalized ACFs of 20 measurements each in the absence and the presence of UV light. It is apparent that the ACFs are bundled into two groups, with the ACFs in the presence of UV light decaying faster, indicating a faster diffusion. Figure 4B displays the recovered D distributions, which shows that the Ds of USN-1 in the presence of UV light are significantly larger than those in the absence of UV light. The mean and 95% confidence intervals are reported in Table 1. The diffusion coefficient was enhanced by a factor of 1.26 (26%). A Student’s t test shows that with a confidence level >99.95%, the diffusions in the presence and absence of UV excitation are different. 8232

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters Scheme 4. Key Portions of the Synthesis of USN-3

substituent in USN-1 is replaced by a tert-butyl group, the motor is reported to have an increased rotational rate from 2 to 3 MHz to >150 MHz.43 Increasing the steric bulk from methyl to tert-butyl likely raises the energy of the intermediate needed for the thermal helix inversion step, making the helix inversion more facile. Likewise, going from the methyl group in USN-1 to the smaller proton in USN-4 will likely lower the energy of the intermediate in USN-4, slowing its rotation. This could account for the slowing of USN-4 relative to USN-1, rather than any effect of unidirectionality vs nondirectional rotation of the rotor. The enhanced diffusion is not caused by the local heating effect of the excitation laser or the UV light. First, both USN-3 and CM-2 serve as excellent control molecules since they have a similar mother-ring structure and the same amount of fluorophores (cy5) as USN-1. However, their diffusion does not increase with UV excitation. Second, we further designed control experiments to exclude the possibility of a heating effect. There are three possible sources for the heating effect: (1) solvent absorption of the excitation laser; (2) fluorophore absorption of the excitation laser; and (3) motor absorption of the UV light. (1) The heating effect caused by the solvent absorption of the excitation laser. It has been well-documented and generally accepted that a mW level laser beam will not cause significant

temperature change in the solvent due to solvent absorption; this has been extensively studied by Hell.44 (2) The heating effect caused by the fluorophore absorption of the 633 nm laser. It is generally accepted in single molecule FCS that the heating caused by fluorophore absorption at the 1.0 mW laser excitation level has a negligible effect upon diffusion. We further confirmed this by varying the 633 nm excitation laser power by a factor of 2.5 (1.2 mW). Note the window for the excitation laser power is very narrow as too much laser power photobleaches the molecules, we do not obtain sufficient signal for too low laser power.45 The corresponding ACF curves and their statistical analyses are shown in Figure S4. The recovered diffusion coefficients using a 3D diffusion model are 0.91 ± 0.11 (× 10−10 m2·s−1) and 0.93 ± 0.10 (× 10−10 m2·s−1) for the 3.0 and 1.2 mW excitation laser powers, respectively. Overlapping of the corresponding ACF curves and their statistical analyses shows that there is no significant difference in diffusion, indicating that there is negligible local heating effect from the 633 nm laser. (3) The heating effect caused by motor absorption of the UV light. The UV light power (10 kW/cm2) is 2 orders of magnitude smaller than that of the 633 nm laser (3.0 MW/ cm2). It is reasonable to infer that the heating caused by the absorption of the UV light is also negligible. However, we noticed that there is a small difference in the UV−vis spectra 8233

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters Scheme 5. Synthesis of USN-4

subsequent rotation of the motors. To further exclude the possibility of the heating effect of the UV light due to this difference in molar absorptivity, we did another control experiment using a previously synthesized nanocar 33. Nanocar 33 has four adamantane wheels and two BODIPY dyes12 whose extinction coefficient is 64,900 M−1 cm−1 at 360 nm (Figure S9). The diffusion coefficient of nanocar 33 was measured in the presence and absence of the UV light illumination on the same confocal fluorescence microscope with a 514 nm laser excitation (0.3 mW or 0.5 MW/cm2). The corresponding ACF curves and their nonlinear least-squares (NLLS) analyses are shown in Figure S10. The recovered diffusion coefficients are 1.11 ± 0.04 (× 10−10 m2·s−1) and 1.10 ± 0.05 (× 10−10 m2·s−1) in the absence and presence of the UV light, respectively. There is no significant difference in diffusion, indicating that the heating effect due to the absorption of the UV light is negligible even for molecules with an absorption coefficient 4× larger at 360 nm. Based on these arguments, we conclude that the observed enhanced diffusion is not due to the heating effect of the excitation UV light or the laser beam. The enhanced diffusion is due to the motor actuation by UV light. The enhanced diffusion for USN-1 and USN-4 molecules can only be observed when the UV photon flux is sufficiently high as our early attempts using low illumination power all failed. At the specified excitation level, the molecule should

Figure 2. Absorption spectra of USN-1, CM-2, USN-3, and USN-4 in ACN.

for USN-1 and its control molecules (Figures S5−S8 and Table S1). Interestingly, the molar absorptivity of the fast rotating motors: USN-1 and USN-4 at 360 nm (15,400 and 14,700 M−1 cm−1, respectively) are larger than those of CM-2 and USN-3 (6400 and 7500 M−1 cm−1, respectively). It is likely that this difference in UV absorption is related to the excitation and 8234

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters

Figure 3. Partial 1H NMR (CD3CN) spectra of half-rotation of the slow motor in 32 and USN-3. (A) Schematic representation of half rotation of the slow motor. (B) Partial 1H NMR spectra of half-rotation of slow motor 32 showing 88% photoisomerization conversion and 99% thermal helix inversion. (C) Partial 1H NMR spectra half-rotation of USN-3 showing 86% photoisomerization conversion and 99% thermal helix inversion. The yields of the conversion were calculated using the integration values of the methyl group (Me).

diffuse by a distance L ∼ 17 nm in the 3D space between two motor actuation events (∼500 ns) according to Einstein eq 1 (D ∼ 1 × 10−10 m2·s−1): L2 = 2nD0t

thiodiethanol (TDE, S(CH2CH2OH)2) was used to mix with ACN to form a binary mixture. 10% of TDE was added so the dynamic viscosity of the solvent was nearly doubled (1.9×), while the viscosity was still low (0.65 mPa s). The diffusion coefficient of USN-1 becomes smaller in the viscous solvent by a factor of 1.7 (Table 2 and Figure 5), qualitatively consistent with Einstein−Stokes equation:

(1)

where L2 is the mean square displacement; n is number of dimensions; D0 is the diffusion coefficient; and t is time interval between two motor excitations. When the UV excitation is close to or over the motor saturation level, t can be approximated as the limiting cycle time of the motor. Under UV activation, eq 1 becomes L2 + r 2 = 2nD′t

D=

kBT 6πηR m

(3)

while USN-1 diffusion was enhanced when the UV excitation was turned on in the viscous solvent, the ratio of the enhancement in the diffusion is approximately constant. As the relative viscosity increased by a factor of 1.9, the diffusion enhancement only changed from 1.26 to 1.23. This shows that the viscosity of the solvent will not significantly affect the diffusion enhancement. In conclusion, we observed that USNs bearing fast lightdriven motors show increased diffusion in the solution phase when the motor is activated by UV light. We demonstrated that the motor rotation is not affected by the fluorophores. Through

(2)

where r is the displacement of the USN after each actuation; D′ is the apparent diffusion coefficient. Note that, r is randomly oriented with respect to L. Thus, an increased D′ by 1.26 times indicates that r, the displacement of the nanomachine under each motor stroke, is ∼8.6 nm, a length several times larger than its molecular size! To investigate how the motor responds to UV light in viscous environments, the diffusion of USNs in a more viscous solution was also investigated. A viscous solvent, 2,2′8235

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters

Figure 4. Comparison of diffusion coefficients of USNs in ACN in the presence and absence of UV light activation. (A, B) USN-1; (C, D) CM-2; (E, F) USN-3; and (G, H) USN-4. (A, C, E, G) are the normalized ACFs of 20 measurements each in the presence and absence of UV light. Red: without UV activation. Blue curves: with UV. (B, D, F, H) are the histograms of recovered diffusion coefficient using nonlinear least-squares fitting from the ACFS. For USN-1 and USN-4, the ACFs are bundled into separate groups in the presence and absence of the UV light, respectively, indicating their diffusion behaviors are significantly different with or without UV light illumination. Using NLLS fitting, the recovered diffusion coefficient Ds of USN-1 and USN-4 in the presence of UV light are significantly larger than those in the absence of UV light (Table 1). The UV light was provided by a gallium indium nitride 365 nm UV LED with an intensity of ∼10 mW. The UV light was optically filtered and tightly focused by a high numerical aperture objective (NA 1.4) to a spot with an estimated diameter of ∼10 μm. The excitation level was ∼1.0 × 104 W cm−2.

Table 1. Apparent Diffusion Coefficients of the USN Series in the Absence and Presence of UV Light Activationa D (no activation) (× 10−10 m2·s−1) USN-1 CM-2 USN-3 USN-4 a

0.92 0.92 0.90 0.89

± ± ± ±

D (UV activation) (× 10−10 m2·s−1)

diffusion coefficient ratio

± ± ± ±

1.26 1.01 1.03 1.10

0.07 0.07 0.06 0.04

1.16 0.93 0.93 0.98

0.10 0.06 0.08 0.04

The diffusion coefficients are reported with 95% confidence intervals using Student’s t-test.

8236

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters

Table 2. Apparent Diffusion Coefficients of USN-1 in Viscous Solutions in the Absence and Presence of UV Light Activationa

a

TDE %

viscosity (mPa·s)

D (no activation) (× 10−10 m2·s−1)

D (UV activation) (× 10−10 m2·s−1)

diffusion enhancement

0 10

0.34 0.65

0.92 ± 0.07 0.53 ± 0.02

1.16 ± 0.10 0.65 ± 0.01

1.26 1.23

The diffusion coefficients are reported with 95% confidence intervals using Student’s t-test.

Figure 5. UV light-enhanced diffusion coefficient of USN-1 molecule in a more viscous solvent (ACN:TDE 9:1). (A) The normalized ACFs in the presence and absence of UV light. Red curves: without UV. Blue curves: with UV activation. (B) Recovered diffusion coefficient distributions.

MHz High Performance Digital NMR. The samples were excited at 365 nm for 1 h using a UVGL-55 lamp (6 W). The yields of the conversion were calculated using the integration values of the methyl group (Me). Sample Preparation for Microscopic Measurements. Cy-5 attached-USN molecules were first dissolved in ACN (Fisher Scientific Inc.) as a stock solution with a concentration of ∼50 μM. In single molecule FCS experiments, the solution was serially diluted in ACN to a final concentration of 2.0 nM. The solution was then sandwiched between a piece of Corning no. 1.5 coverglass and a piece of glass slide using two pieces of double-sided Scotch tape (∼90 μm) as the spacers. Finger nail polish was used to seal the solution in the chamber. To study the viscosity effect on the increased diffusion by UV-light, 2,2′thiodiethanol (TDE, Sigma-Aldrich) was used to form a binary mixture with ACN at different compositions. All solutions were prepared fresh daily. Confocal Single Molecule Fluorescence Correlation Spectroscopy with UV Activation. The excitation was provided by an unpolarized 633 nm HeNe laser focused to the diffraction limited spot with an output power of (∼3.0 MW/cm2) (Uniphase) unless otherwise specified. The excitation beam was collimated to overfill the back aperture of a microscope objective (Nikon, 100× Plan Apo/1.40−0.7 oil-immersed). The fluorescence signal was filtered through a 655 long-pass dichroic mirror and a 684 ± 24 nm band-pass filter and imaged into a piece of multimode fiberoptics (Thorlabs) and detected by an avalanche photodiode (PerkinElmer, SPCMAQRH-15-FC). The diameter of the fiberoptics was 50 μm (∼0.8 AU). A programmable counting board was used for photon counting. In the UV activation experiments, a gallium indium nitride UV LED emitting at 365 nm was used. The LED emission was filtered using a 350 ± 25 nm optical filter and focused by an oil immersion objective (NA 1.4) from the opposite side of the microscope objective. The total power of the UV light was ∼10 mW after optical filter cleaning. The UV spot size was estimated to ∼10 μm. The UV activation and no activation experiments were always collected in pairs using the same solution and at the same collection spot. The sequence of

careful design of control molecules with no motor, a slow motor, or a non-unidirectionally rotating motor, we found that a fast unidirectional rotating motor at the MHz range is crucial for increased diffusion, but a non-unidirectional motor can also work, albeit less effectively. No significant change in the diffusion enhancement ratio with increased solvent viscosity was observed. The enhancement of 26% in diffusion suggests that the USN molecules will give ∼9 nm step upon each motor actuation. While the mechanism of movement is still under study, the activated motion of the molecular-sized entities is possible in spite of Brownian motion in solution. This study provides insight in molecular designs for submersible nanomachines. Methods. General Synthetic Methods. 1H NMR and 13C NMR spectra were recorded at 400, 500, or 600 and 100, 125, or 150 MHz, respectively. Chemical shifts (δ) are reported in ppm from tetramethylsilane (TMS). FTIR spectra were recorded using a FTIR infrared microscope with ATR objective with 2 cm−1 resolution. All glassware was oven-dried overnight prior to use. Reagent grade tetrahydrofuran (THF) and ether (Et2O) was distilled from sodium benzophenoneketyl under N2 atmosphere. Triethylamine (NEt 3 ), dichloromethane (CH2Cl2), and N,N′-dimethylforamide (DMF) were distilled from calcium hydride (CaH2) under N2 atmosphere. THF and NEt3 were degassed with a stream of argon for 15 min before being used in the Sonogashira coupling reactions. All palladium-catalyzed reactions were carried out under argon atmosphere, while other reactions were performed under N2 unless otherwise noted. All other chemicals were purchased from commercial suppliers and used without further purification. Flash column chromatography was performed using 230−400 mesh silica gel from EM Science. Thin-layer chromatography was performed using glass plates precoated with silica gel 40 F254 0.25 mm layer thickness purchased from EM Science. UV−vis Measurements. UV−vis spectra were recorded on a Shimadzu UV-2450 or a HP 8543 UV−vis spectrophotometer using spectroscopic grade acetonitrile. Monitoring of Half Rotation of the Motor. The 1H NMR spectra of 1 mM solutions of slow motor 32 and USN-3 in CD3CN were recorded using a Bruker AVANDE III HD 600 8237

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters

(5) Shirai, Y.; Osgood, A. J.; Zhao, Y.; Yao, Y.; Saudan, L.; Yang, H.; Yu-Huang, C.; Alemany, L. B.; Sasaki, T.; Morin, J.-F.; Guererro, J. M.; Kelly, K. F.; Tour, J. M. J. Am. Chem. Soc. 2006, 128, 4854−4864. (6) Shirai, Y.; Osgood, A. J.; Zhao, Y. M.; Kelly, K. F.; Tour, J. M. Nano Lett. 2005, 5, 2330−2334. (7) Chiang, P. T.; Mielke, J.; Godoy, J.; Guerrero, J. M.; Alemany, L. B.; Villagómez, C. J.; Saywell, A.; Grill, L.; Tour, J. M. ACS Nano 2012, 6, 592−597. (8) Claytor, K.; Khatua, S.; Guerrero, J. M.; Tour, J. M.; Link, S. J. Chem. Phys. 2009, 130, 164710. (9) Khatua, S.; Guerrero, J. M.; Claytor, K.; Vives, G.; Kolomeisky, A. B.; Tour, J. M.; Link, S. ACS Nano 2009, 3, 351−356. (10) Khatua, S.; Godoy, J.; Tour, J. M.; Link, S. J. Phys. Chem. Lett. 2010, 1, 3288. (11) Vives, G.; Guerrero, J. M.; Godoy, J.; Khatua, S.; Wang, Y.-P.; Kiappes, J. L.; Link, S.; Tour, J. M. J. Org. Chem. 2010, 75, 6631−6643. (12) Chu, P.-L. E.; Wang, L.-Y.; Khatua, S.; Kolomeisky, A. B.; Link, S.; Tour, J. M. ACS Nano 2013, 7 (1), 35−41. (13) Fournier-Bidoz, S.; Arsenault, A. C.; Manners, I.; Ozin, G. A. Chem. Commun. 2005, 4, 441−443. (14) Paxton, W. F.; Kistler, K. C.; Olmeda, C. C.; Sen, A.; St. Angelo, S. K.; Cao, Y.; Mallouk, T. E.; Lammert, P. E.; Crespi, V. H. J. Am. Chem. Soc. 2004, 126, 13424−13431. (15) Solovev, A. A.; Mei, Y.; Bermúdez Ureña, E.; Huang, G.; Schmidt, O. G. Small 2009, 5, 1688−1692. (16) Howse, J. R.; Jones, R. A. L.; Ryan, A. J.; Gough, T.; Vafabakhsh, R.; Golestanian, R. Phys. Rev. Lett. 2007, 99, 048102. (17) Pavlick, R. A.; Sengupta, S.; McFadden, T.; Zhang, H.; Sen, A. Angew. Chem., Int. Ed. 2011, 50, 9374−9377. (18) Muddana, H. S.; Sengupta, S.; Mallouk, T.; Sen, A.; Butler, P. J. J. Am. Chem. Soc. 2010, 132, 2110−2111. (19) Sengupta, S.; Dey, K. K.; Muddana, H. S.; Tabouillot, T.; Ibele, M. E.; Butler, P. J.; Sen, A. J. Am. Chem. Soc. 2013, 135, 1406−1414. (20) McDermott, J. J.; Kar, A.; Daher, M.; Klara, S.; Wang, G.; Sen, A.; Velegol, D. Langmuir 2012, 28, 15491−15497. (21) Duan, W.; Ibele, M.; Liu, R.; Sen, A. Eur. Phys. J. E: Soft Matter Biol. Phys. 2012, 35, 77. (22) Hong, Y.; Diaz, M.; Córdova-Figueroa, U. M.; Sen, A. Adv. Funct. Mater. 2010, 20, 1568−1576. (23) Giudicatti, S.; Marz, S. M.; Soler, L.; Madani, A.; Jorgensen, M. R.; Sanchez, S.; Schmidt, O. G. J. Mater. Chem. C 2014, 2, 5892−5901. (24) Pavlick, R. A.; Dey, K. K.; Sirjoosingh, A.; Benesi, A.; Sen, A. Nanoscale 2013, 5, 1301−1304. (25) Godoy, J.; García-López, V.; Wang, L.-Y.; Rondeau-Gagne, S.; Marti, A.; Link, S.; Tour, J. M. Tetrahedron 2015, 71, 5965−5972. (26) Klok, M.; Boyle, N.; Pryce, M. T.; Meetsma, A.; Browne, W. R.; Feringa, B. L. J. Am. Chem. Soc. 2008, 130, 10484−10485. (27) Koumura, N.; Geertsema, E. M.; van Gelder, M. B.; Meetsma, A.; Feringa, B. L. J. Am. Chem. Soc. 2002, 124, 5037−5051. (28) Chen, J.; Kistemaker, J. C.; Robertus, J.; Feringa, B. L. J. Am. Chem. Soc. 2014, 136, 14924−14932. (29) Taylor, G. I. Proc. R. Soc. London, Ser. A 1951, 209, 447−461. (30) Ludwig, W. Zeit. F. Vergl. Physiol. 1930, 13, 397−504. (31) Purcell, E. M. Am. J. Phys. 1977, 45, 3−11. (32) Happel, J., Brenner, H. Low Reynolds Number Hydrodynamics; Prentice-Hall: Englewood Cliffs, NJ, 1965. (33) Happel, J., Brenner, H. Low Reynolds Number Hydrodynamics, 2nd ed.; Springer: New York, 1981. (34) Nelson, B. J.; Kaliakatsos, I. K.; Abbott, J. J. Annu. Rev. Biomed. Eng. 2010, 12, 55−85. (35) Zhang, L.; Abbott, J. J.; Dong, L.; Peyer, K. E.; Kratochvil, B. E.; Haixin, Z.; Bergeles, C.; Nelson, B. J. Nano Lett. 2009, 9, 3663−3667. (36) Peyer, K. E.; Tottori, S.; Qiu, F.; Zhang, L.; Nelson, B. J. Chem. Eur. J. 2013, 19, 28−38. (37) Lauga, E.; Powers, T. R. Rep. Prog. Phys. 2009, 72, 96601− 96637. (38) Wang, J. Nanomachines. Fundamentals and Applications. WileyVCH: Weinheim, 2013; pp 13−34.

collection has no observable effect on the diffusion coefficient measurements. The integration time was 30−100 μs, depending on the diffusion speed of the USN molecules. The acquired data were analyzed using MATLAB and Origin software. Data Analysis. When a molecule diffuses into the detection volume of a confocal fluorescence microscope, a photon burst will be generated and recorded. A typical fluorescence intensity trace for USN molecules diffusing in ACN is shown in Figure 2. The ACF of the intensity trace follows a 3D model eq 4:32 G (τ ) =

1 1 N 1 + τ /τdiff

1 1 + τ /S2τdiff

(4)

where ⟨N⟩ is the average number of emitters in the probe volume; S is the aspect ratio of the probe volume rz/rxy; τdiff is the characteristic diffusion time assuming that the emitter has an isotropic diffusion coefficient D, eq 5:

τdiff =

2 rxy

4D

(5)

where rxy and rz are the distances from the center to where the emission intensity drops to 1/e2 in the lateral and axial directions. The rxy and rz were estimated to be ∼300 and ∼900 nm, respectively. The apparent diffusion coefficient D in the absence and presence of UV activation was obtained through NLLS fitting of the experimentally acquired data.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.nanolett.5b03764. Detailed experimental procedures and data (PDF) Additional spectroscopic data (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: *E-mail: *E-mail: *E-mail:

[email protected]. [email protected]. [email protected]. [email protected].

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS G.W. acknowledges North Carolina State University start-up funds and FRPD Award. A.B.K. acknowledges support from the National Science Foundation (CHE-1360979), the National Institutes of Health (1R01GM094489-01) and the Welch Foundation (C-1559). J.M.T. and A.A.M. acknowledge support from the National Science Foundation (CHE-1007483). The authors thank Prof. Felix Castelleno for his insightful discussions.



REFERENCES

(1) Tour, J. M. Chem. Mater. 2014, 26, 163−171. (2) Alberts, B. Cell 1998, 92, 291−294. (3) Kinbara, K.; Aida, T. Chem. Rev. 2005, 105, 1377−1400. (4) Shirai, Y.; Morin, J. F.; Sasaki, T.; Guerrero, J. M.; Tour, J. M. Chem. Soc. Rev. 2006, 35, 1043−1055. 8238

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Letter

Nano Letters (39) Morin, J.-F.; Shirai, Y.; Tour, J. M. Org. Lett. 2006, 8, 1713− 1716. (40) Neupane, B.; Chen, F.; Sun, W.; Chiu, D. T.; Wang, G. F. Rev. Sci. Instrum. 2013, 84, 043701. (41) Enderlein, J.; Gregor, I.; Patra, D.; Fitter. Curr. Pharm. Biotechnol. 2004, 5, 155−161. (42) Zhong, Z. M.; Lowry, M.; Wang, G. F.; Geng, L. Anal. Chem. 2005, 77, 2303−2310. (43) Klok, M. Motors for use in Molecular Nanotechnology. Ph.D. Thesis, University of Groningen, 2009. http://dissertations.ub.rug.nl/ faculties/science/2009/m.klok/ (accessed October 24, 2015). (44) Schönle, A.; Hell, S. W. Opt. Lett. 1998, 23, 325−327. (45) Mukhopadhyay, A.; Zhao, J.; Bae, S. C.; Granick, S. Rev. Sci. Instrum. 2003, 74, 3067.

8239

DOI: 10.1021/acs.nanolett.5b03764 Nano Lett. 2015, 15, 8229−8239

Unimolecular Submersible Nanomachines. Synthesis, Actuation, and Monitoring.

Unimolecular submersible nanomachines (USNs) bearing light-driven motors and fluorophores are synthesized. NMR experiments demonstrate that the rotati...
NAN Sizes 1 Downloads 7 Views